Download Ion Exchange Resins: Catalyst Recovery and Recycle

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Metal carbonyl wikipedia , lookup

Discodermolide wikipedia , lookup

Alkene wikipedia , lookup

Elias James Corey wikipedia , lookup

Cracking (chemistry) wikipedia , lookup

Physical organic chemistry wikipedia , lookup

Kinetic resolution wikipedia , lookup

Petasis reaction wikipedia , lookup

Strychnine total synthesis wikipedia , lookup

Asymmetric induction wikipedia , lookup

Polythiophene wikipedia , lookup

Baylis–Hillman reaction wikipedia , lookup

Ene reaction wikipedia , lookup

Fischer–Tropsch process wikipedia , lookup

Enantioselective synthesis wikipedia , lookup

Asymmetric hydrogenation wikipedia , lookup

Ring-closing metathesis wikipedia , lookup

Hydrogenation wikipedia , lookup

Hydroformylation wikipedia , lookup

Transcript
Subscriber access provided by UNIV SCI & TECH OF CHINA
Review
Ion Exchange Resins: Catalyst Recovery and Recycle
Pierluigi Barbaro, and Francesca Liguori
Chem. Rev., 2009, 109 (2), 515-529• DOI: 10.1021/cr800404j • Publication Date (Web): 23 December 2008
Downloaded from http://pubs.acs.org on March 5, 2009
More About This Article
Additional resources and features associated with this article are available within the HTML version:
•
•
•
•
Supporting Information
Access to high resolution figures
Links to articles and content related to this article
Copyright permission to reproduce figures and/or text from this article
Chemical Reviews is published by the American Chemical Society. 1155 Sixteenth
Street N.W., Washington, DC 20036
Chem. Rev. 2009, 109, 515–529
515
Ion Exchange Resins: Catalyst Recovery and Recycle
Pierluigi Barbaro* and Francesca Liguori
Istituto di Chimica dei Composti Organo Metallici - Consiglio Nazionale delle Ricerche, Area di Ricerca di Firenze, Via Madonna del Piano 10,
50019 Sesto Fiorentino, Firenze, Italy
Received June 9, 2008
Contents
1. Introduction
515
2. Ion-Exchange Resins and Homogeneous Catalysts 516
Immobilization
3. Applications
517
3.1. Carbonylations
517
3.2. Hydroformylations
518
3.3. Hydrogenations
519
3.3.1. Asymmetric Hydrogenations
520
3.4. Oxidations
521
3.4.1. Alkene Epoxidations
521
3.4.2. Asymmetric Dihydroxylations
522
3.4.3. Various Oxidations
523
3.5. Polymerizations
524
3.6. Miscellaneous
524
4. Conclusions
525
5. Acknowledgments
526
6. References
526
1. Introduction
Chemical synthesis is greatly facilitated by catalysis and
further on by catalyst recovery and recycle. Catalyst reuse
increases the overall productivity and cost effectiveness of
chemical transformations while minimizing their environmental impact, ultimately contributing considerably to the
sustainability of chemical processes. Indeed, catalyst recycle
fits most “principles of green chemistry”1-4 and was included
as a priority in several strategic research agendas both in
the United States and in Europe.5,6
To date, catalytic technologies leading to fine-chemicals
production, and particularly in selective, large-scale processes, are still largely dominated by homogeneous catalysts,7-13
whose separation from the reaction products and reuse is a
major concern.14-16
Due to the easier workup and integration in reactor
equipments, the chemical industry has a strong preference
for solid catalysts, which, however, usually do not provide
selectivities (chemo, regio, or stereo) comparable to those
observed in the homogeneous phase.17,18
There is therefore a clear need to develop new concepts
bridging heterogeneous and homogeneous catalysis and to
apply these to the engineering of catalytic devices for the
industrial production of fine chemicals.19-25
In order to easily recover and recycle homogeneous
catalysts, various techniques were developed over the last
* To whom correspondence
[email protected].
should
be
addressed.
E-mail:
20 years involving the immobilization of a catalyst precursor
onto an insoluble support material, so that the catalyst can
be quantitatively separated by filtration and recycled. This
topic was extensively and excellently reviewed in the
past.26-39 A variety of solids, often highly sophisticated, have
been exploited for this purpose, including inorganic (silica,
clays, zeolites, metal oxides, heteropolyacids, etc.),40-42
organic (carbon, dendrimers, polymeric ligands, polyelectrolytes, etc.),43-45 and hybrid materials.46-48 The active
species can be immobilized on these supports by covalent
or noncovalent binding, i.e. by adsorption, electrostatic
interaction, or entrapment.49-52
Preformed, molecular homogeneous chemical catalysts
(usually metal complexes or organometallic compounds) are
most conveniently anchored to diverse materials through
noncoValent binding. This approach, hereinafter referred to
as heterogenization of homogeneous catalysts,53-57 has
several benefits:
-there is no need for the chemical modification, neither of
the support nor of the catalyst, usually required using a
covalent binding strategy,
-catalysts of known activity and selectivity are used,
-the anchoring procedures are simple,
-the problems arising from metal loading are minimized,
-the catalyst’s active sites can be easily characterized,
-a systematic design of new, predictable catalysts is enabled.
The catalysts thus obtained can be classified into the more
general family of single-site catalysts whose features were
detailed by Thomas et al.58 The catalytic performances of
these heterogenized catalysts may vary enormously depending on the immobilization method and on the support, so
that it is of outmost importance to get a systematic picture
of favorable and unfavorable factors and to test different
support materials.
This paper reviews the recent achievements in the field
of the heterogenization of homogeneous chemical catalysts
onto ion-exchange resins. The main focus will be on those
applications for which the recycling of the catalyst has been
demonstrated. This matter has been partially reviewed up to
the end of 2004.59-65 The present manuscript coVers the most
significant papers that appeared in the literature from
January 2005 to April 2008. Some additional aspects are
also taken into account, including reference to previous but
unreviewed works. Applications will be described according
to the reaction involved, irrespective of the nature of the
supported metal or the type of the resin used.
The immobilization of biological catalysts,66,67 the use of
ion-exchange resins as Brönsted and/or Lewis solid acid
catalysts (alone or in combination with metal ions),68-71 or
their use in simple extraction procedures (e.g., for metal
10.1021/cr800404j CCC: $71.50  2009 American Chemical Society
Published on Web 12/23/2008
516 Chemical Reviews, 2009, Vol. 109, No. 2
Barbaro and Liguori
Scheme 1
Pierluigi Barbaro was born in Firenze in 1962. He obtained his Ph.D. in
chemistry from the University of Florence and completed his postdoctoral
work at ETH in Zurich, later becoming a permanent researcher at the
Istituto di Chimica dei Composti Organo Metallici - Consiglio Nazionale
delle Ricerche, Italy. He has been a member of the Network of Excellence
IDECAT since 2005, and he is coordinator of the Initial Training Network
NANO-HOST, funded by the European Commission under the VII° FP.
In 1991 he was awarded with the national prize “Federchimica“. He is
the author of more than 70 research papers and 6 patents and is the
editor of a book. His main research interest is in the field of homogeneous,
heterogeneous, and asymmetric catalysis, with focus on homogeneous
supported catalysts and nanostructured catalysts for sustainable production
processes, microscopy, and nuclear magnetic resonance.
Francesca Liguori (born in 1971, Prato, Italy) received her degree in
Organic Chemistry in 1999 at the University of Florence, Italy, with a
thesis on the synthesis of heterocyclic alkaloids. After some postdegree
research collaborations at the Department of Organic Chemistry of the
University of Florence, she received her Ph.D. in Chemistry in 2005 with
a thesis on the synthesis of gangliosides, mimetics of antitumor antigens.
In 2006 she joined the Institute ISMAC-CNR (Milan, Italy) as a postdoctoral
fellow, working on the synthesis of ethylene based copolymers with antioxidant activity. Since 2007 she has been a grant-holder at the Institute
ICCOM-CNR (Florence, Italy), with a project funded by the Network of
Excellence IDECAT on heterogeneous catalysts for asymmetric hydrogenation reactions. Her research interest extends over a wide range of
organic product chemistry, heterocycles, carbohydrates, and organometallic
complexes and, more recently, to the synthesis and reactivity studies of
homogeneous catalysts heterogenized on diverse materials.
catalysts recovery),72,73 as well as their use in the pyrolytic
preparation of carbon-based catalysts,74,75 will not be
considered.
2. Ion-Exchange Resins and Homogeneous
Catalysts Immobilization
A systematic description of ion-exchange resins is out of
the scope of this paper. Comprehensive classifications and
overviews of their properties were published elsewhere.76-78
A short mention of the resin features which may affect the
efficiency of the supported homogeneous catalysts and the
viability of their recycle is provided here, aimed at a critical
coverage of the cited literature.
Common uses of ion-exchange resins include water
purification, metal recovery and separation, ion substitution,
acid-base catalysis, as sensors or as solid electrolytes (e.g.,
in fuel cells, electrolyzers, electrodialysis devices), in chemical, food, and beverage industries, and in power, nuclear,
semiconductor, and pharmaceutical industries.79-81
Most ion-exchange resins are based on cross-linked
polystyrene-divinylbenzene copolymers bearing ion-exchanging functional groups (Scheme 1).82,83 Classification
of resins usually involves four main groups:
Cation exchanger (with anionic functionalities and positively charged mobile ions)
-strong acid exchange (e.g., containing sulfonic acid groups
or the corresponding salts)
-weak acid exchange (e.g., containing carboxylic acid groups
or the corresponding salts)
Anion exchanger (with cationic functionalities)
-strong base exchange (e.g., containing quaternary ammonium groups)
-weak base exchange (e.g., containing ammonium groups)
Other ion-exchanging materials include homopolystyrene
and acrylic based resins and Nafion, a perfluorinated polymer
containing sulfonic acid heads, whose structure and properties
were previously carefully reviewed.84-88
Cross-linkage, typically from 0.5 to 20%, controls the resin
porosity.89 Low cross-linked resins have a gel (microporous)
structure, whereas higher cross-linkage degrees result in
macroreticular (macroporous) resins. Porosity affects some
bulk properties of the resins which have consequences on
their catalytic applications, i.e. swelling, capacity, equilibration rate, and selectivity. Usually, the lower the cross-linking
percentage, the higher the moisture content, the equilibration
rate, the loading capacity (typically from 1.5 to 10 mequiv/g
on a dry basis), and the ability to accommodate larger ions.
Non-cross-linked polymers are rarely used due to their lower
stability (chemical, mechanical, thermal).
Swelling of the resins in the solvent of use is crucial for
their behavior. Swelling volumes were carefully investigated,
and increments up to 800% upon decreasing the cross-linking
percentage were frequently found.90-92 Hence, gel type resins
are generally preferred over macroporous ones due to
enhanced mass transfer inside the polymer beads, resulting
in good active-sites accessibility to all soluble reactants.93
Ion-exchange resins are commercial products commonly
available in the form of small beads (16 ÷ 400 mesh, 1180
÷ 38 µm diameter) or as membranes. Shape and size allow
Ion Exchange Resins
Chemical Reviews, 2009, Vol. 109, No. 2 517
Scheme 2
these materials to be easily and quantitatively recovered by
simple filtration or decantation. This is not the case for other
powdered materials often used as catalysts support (e.g.,
silica, zeolites, carbon, etc.). When the particle size is about
1 µm or less, they might not settle in the solution within a
short time, and it would be very difficult to collect them for
recycling. The separation of the catalyst thus requires
centrifugation or ultrafiltration. Very fine powders may also
clog or poison the reactors or the autoclaves employed in
the catalytic experiments.
Immobilization of preformed, charged metal complexes
onto ion-exchange resin is an equilibrium process driven by
noncovalent electrostatic interactions (Scheme 2, strong
cation exchanger example). The affinity and selectivity of
resins varies with the ionic size and charge of the ions.
Generally, the affinity is greatest for large ions with high
valence.94-96 Compared to other insoluble support materials,
ion-exchange resins usually offer considerable advantages
as catalysts carriers,97 as summarized in Table 1.
The engineering of heterogenized catalyst should be
accomplished with minimal chemical manipulations and
energy consumption.98 Ion-exchange resins are perfectly
located in this setting: catalysts which are easy to use and
to recover are readily obtainable through them. Reproducibility and ease of characterization is also ensured, which is
not obvious for conventional heterogeneous catalysts. Performances, both activities and selectivities, comparable with
those of the homogeneous counterparts are observable as
well.
Batch or continuous flow operations with fixed beds of
immobilized catalysts should be possible with minimum
leaching of active species or metal impurities into the reaction
solvent, particularly as far as the production of pharmaceutical precursors is concerned.99,100 This can be an issue with
noncovalently bound heterogenized catalysts, including ionexchange resins. Nonetheless, ion-pair formation is sufficiently strong in exchange resins to allow for the uncommon
advantage that the solvent in which the homogeneous
catalysts is soluble can be used as the reaction medium.
Viability of the use of water as solvent is also worthy to be
Figure 1. Proposed cycle for the rhodium catalyzed methanol
carbonylation.
underlined due to its environmental implications.101,102 The
combination of the optimal resin swelling with the usual lowsolubility of organometallic complexes in this medium may
result in efficient catalysts and in reduced leaching.
3. Applications
3.1. Carbonylations
One of the most important examples of carbonylation
reactions is the manufacture of acetic acid. More than 60%
of the current industrial capacity of acetic acid is based on
the so-called “Monsanto” methanol carbonylation process
developed in the late 1960s.103 The process is based on a
well established homogeneous iodide/rhodium-catalyzed
system operating under mild conditions (150-220 °C, 30-40
atm CO) and exhibiting high selectivity to acetic acid (99%
yield from methanol).104 Several mechanistic studies provided
a detailed description of the catalytic cycle (Figure 1).105
The system requires a substantial quantity of water
(14-15% w/w) to achieve high catalyst activity and maintain
good catalyst stability because the hydration of acetyl iodide
in the presence of excess water gives acetic acid and
hydrogen iodide to complete the cycle. As a consequence,
the homogeneous system is affected by some important
drawbacks in relation to the solubility of the catalysts, the
loss of the expensive rhodium metal due to precipitation in
the purification steps, and the high corrosion rates caused
by iodide.
Accordingly, several studies were reported, aimed at the
heterogenization of rhodium catalysts for methanol carbo-
Table 1. Ion-Exchange Resins and Their Use as Support for Catalyst Immobilization
advantages
commercially available in several varieties
low-cost
reasonably stable
defined amounts of anchoring sites
easy immobilization procedure
ease of handling
particle size allows quantitative recovery by simple filtration
catalyst efficiency comparable to homogeneous reactions
ease of catalyst recycle
negligible to low metal leaching
compatible with many reaction solvents, including water
adaptable to the engineering of reactors
disadvantages
reaction rates may decrease upon recycle
leaching sometimes not acceptable for industrial applications
518 Chemical Reviews, 2009, Vol. 109, No. 2
Scheme 3
nylation on diverse solid supports.106 The fact that all
rhodium complexes involved in the catalytic cycle are anionic
suggested this system to be an excellent candidate for
noncovalent, ionic immobilization,107 and, indeed, this was
one of the earliest applications of molecular catalysts
supported on ion exchange resins.108,109 Most importantly,
Chiyoda/UOP developed an industrial process, the Acetica
Process, based on this technology.110-112 The actual process
employs beads of 33% w/w cross-linked 4-vinylpyridine/
divinylbenzene copolymer resin (macroporous, Reillex 425,
Reilly Corporation) as solid support.113 Treatment of the resin
with RhCl3 and a methyl iodide/methanol/acetic acid mixture
under reaction conditions (190 °C, 5.0 MPa carbon monoxide
pressure, low water content) generates in situ both the
N-methylpyridinium iodide ion-exchanger derivative and,
subsequently, the immobilized catalyst, by substitution of
iodide with the anionic rhodium(I) precursor [Rh(CO)2I2]-,
which forms under these conditions (Scheme 3).114 The
typical Rh loading observed for the heterogeneous catalyst
is 0.8% w/w.
Investigations carried out using infrared spectroscopy and
EXAFS measurements (extended X-ray absorption fine
structure) indicated that the complex [Rh(CO)2I2]- supported
on quaternized poly(4-vinylpyridine-co-styrene-co-divinylbenzene) resins adopts an identical structure to that found
in solution,115 thus confirming the presence of the same active
species in the homogeneous and heterogeneous processes.108
This study also revealed that the immobilized [Rh(CO)2I2]complex undergoes analogous reactivity and kinetics with
CH3I as in homogeneous phase, except for a subtle cation
effect on the rate of the addition, being lower for the resinsupported complex and highest for soluble, monomeric
N-methylpyridinium salts. However, close similarities in the
rate constants suggested that the permeability of the resin to
CH3I is high and no mass-transfer limitations arise in the
heterogeneous phase.
Recent studies on Reillex 425 showed that this resin
markedly increased its swellability in water/methanol/acetic
acid mixtures after iodomethylation.116 The combination of
SEM (scanning electron microscopy), ISEC (inverse steric
exclusion chromatography), and ESR (electron spin resonance) techniques indicated this material develops a well
defined nanoporosity and a good molecular accessibility in
the swollen state that may help rationalize the catalytic
efficiency of the ionically immobilized rhodium precursor.
The use of an heterogenized catalyst in the AceticaTM
Process displays several advantages over the conventional
homogeneous-phase system:117-119
Barbaro and Liguori
Scheme 4
-high reactor productivity can be obtained because high
catalyst concentrations are achievable without solubility
limitations;
-rhodium loss is significantly reduced, as no catalyst separation from the reaction liquid is needed;
-the lower water content required (3-7%) minimizes byproduct formation (propionic acid and CO2 Via the water gas shift
reaction) and corrosion, due to reduced hydrogen iodide
content;
-tolerance of the resin to elevated temperatures and pressures
provides very stable catalysts featuring no deactivation after
prolonged continuous operations.
3.2. Hydroformylations
The catalytic homogeneous hydroformylation of alkenes
to aldehydes is one of the largest industrial processes,
accounting for a production of more than 6 × 106 tons per
year as plasticizers, detergents, and fragrances (Scheme 4).120
Heterogenization of soluble hydroformylation catalyst,
aimed at precious metals recovery and facile product
separation, was attempted using a variety of materials,121
including ion-exchange resins.122 Hydroformylation of 1-hexene was recently accomplished using the known TPPTSRh(I) catalyst,123 after immobilization onto macroporous
Amberlite IRA-93 anion exchange resin (TPPTS ) sodium
triphenylphosphine trisulfonate).124 The heterogeneous catalyst was prepared by derivatization of the resin with the
phosphine ligand by anion exchange, followed by rhodium
complexation (Scheme 5). The final rhodium content in the
dry resin was ca. 0.2% (w/w).
The solvent used in the hydroformylation proved to be
extremely important in determining the selectivity and the
rate of the reaction. Several solvents were screened for
conversion, hexene isomerization, selectivity to aldehyde, and
n/iso ratio, with the former two being higher for benzene
(99.3% conversion, 53.4% isomerization) and cyclohexene
(99.0%, 58.7%), and the latter two for ethanol (71.4%
selectivity, 2.31 n/iso ratio). Although no explanations were
given for the observed solvent effect, a kinetic study carried
out in ethanol revealed the reaction to be first order in
hydrogen partial pressure and 1-hexene concentration with
a maximum in the rate depending on the carbon monoxide
partial pressure.
Interestingly, unlike the case of the homogeneous phase
reaction, the n/iso-aldehyde ratio remained constant upon
reaction progress. This behavior was explained in terms of
hindered access of the isomerized olefin to the resin bound
catalyst, which, thus, difficultly undergoes hydroformylation
to the branched aldehyde. The performance of the heterogeneous catalyst was also checked on recycling. No major
changes were observed in activity, selectivity, n/iso ratio,
and rhodium leaching (below 0.1 ppm), over five catalysis
Ion Exchange Resins
Chemical Reviews, 2009, Vol. 109, No. 2 519
Scheme 5
Scheme 6
cycles. Turnover frequencies comparable to those found for
comparable homogeneous systems were observed throughout
(ca. 3000 h-1).121b
3.3. Hydrogenations
The catalytic hydrogenation of CdC and CdO bonds is
among the most useful tool for the production of fine
chemicals, including pharmaceuticals, agrochemicals, and
fragrances.125-127 Lately, alkynes, cinnamaldehyde, and acetophenone were hydrogenated using the well-known watersoluble ruthenium(II) and rhodium(I) complexes
[RuCl2(mtppms)2]2 and RhCl(mtppms)3,128-130 after immobilization on commercially available anion-exchange
resins (mtppms ) [m-sulfonatopheny]diphenylphosphine
sodium salt).131 The DEAE-Molselect (strong, diethylaminoethyl groups), QAE-Sephadex (strong, diethyl(2-hydroxypropyl)aminoethyl groups), and Lewatit MonoPlus (strong,
quaternary amine) exchangers were used as supports.
The ruthenium derivative showed to hydrogenate diphenylacetylene to cis-stilbene with good selectivity under mild
reaction conditions (toluene/ethanol ) 1/1, 30-80 °C, 30
bar H2) (Scheme 6). The conversion increased by increasing
the temperature, albeit with a decrease in selectivity. At the
optimal temperature (50 °C), up to 83% selectivity was
observed with an efficiency higher than that of the analogous
water/toluene biphasic system (TOF ) 25.1 h-1 heterogeneous, 1.96 h-1 biphasic).132a The catalyst was reused up to
20 times with only a slight decrease in activity and
selectivity. No data were reported on possible metal leaching,
however.
Consistently with the known full-hydrogenation attitude
of rhodium, the Rh catalyst was less selective compared to
ruthenium in diphenylacetylene reduction, providing 1,2diphenylethane in 62% yield, at 87.5% total conversion (50
°C, 30 bar H2).
As for the hydrogenation of R,β-unsaturated carbonyl
compounds by molecular hydrogen, this can be efficiently
accomplished using transition metal catalysts, although
chemoselectivity may be an issue.133 The hydrogenation of
the CdC bond, in fact, generally competes favorably with
that of the CdO group.134
[RuCl2(mtppms)2]2 immobilized on Lewatit MonoPlus
selectiVely hydrogenated trans-cinnamaldehyde (1) to 3-phenylpropanal (2), in low conversions (18% at 60 °C, 80 bar
H2, ethanol, TOF ) 5.2 h-1) (Scheme 7). Interestingly, this
result compares positively with the selectivity provided by
both parent homogeneous and biphasic systems.132b The
Scheme 7
homogeneous phase RuCl2(PPh3)3/toluene catalyst gave
3-phenyl-1-propanol (4) in 42% yield, with a selectivity to
the saturated aldehyde of 23%, under comparable conditions
(100 °C, 30 atm H2, TOF (2) ) 6.3 h-1). By contrast,
cinnamyl alcohol (3) was the major product (83%) using the
1 /1 ) water/toluene [RuCl2(mtppms)2]2 two-phase system
(100 °C, 30 atm H2, TOF (2) ) 1.2 h-1), with a selectivity
to 2 of 15%.135
Previous studies carried out on the analogous chlorobenzene/water-phosphate buffer biphasic system demonstrated
the selectivity of the reaction to be ruled by the relative
concentrations of the catalytically active species, either the
monomeric[H2Ru(mtppms)4]or[HRuCl(mtppms)3]complex.132c-e
The former, which is the predominant ruthenium species at
pH values greater than 6, led to the preferential production
of alcohol 3 (up to 95% selectivity) Via the classical Ruassisted hydride transfer mechanism to the carbonyl
group.132e,f,136 The latter, which dominates below pH 5, gave
aldehyde 2 as the main product (ca. 60%), Via dissociation
of a phosphine ligand followed by coordination and hydrogenation of the CdC bond (Figure 2).132b,e,g,h
Although a rationale for the selectivity observed using ionexchange supported tppms catalysis has not been provided,
the above findings demonstrate a strong influence of the ionic
solid support on the performance of the immobilized molecular catalysts and suggest that the effect on the selectivity
can be attributed to the subtle role played by the support in
the stabilization of intermediates similar to those sketched
in Figure 2, maybe as a consequence of favorable dipolar
interactions of the CdO group with the support or of the
reduced steric hindrance of these intermediates.
It must be noted that the best catalysts to date, both in
term of efficiency and (stereo) chemoselectivity, in the
reduction of the CdO bond in R,β-unsaturated carbonyl
compounds is the homogeneous, bifuctional system developed by Noyori based on [RuCl2(phosphane)2(1,2-diamine)]
and an alkaline base, which produces cinnamyl alcohol from
cinnamaldehyde in 99.9% selectivity.137
Figure 2. Proposed intermediates for the catalytic hydrogenation
of the CdC bond in R,β-unsaturated aldehydes. See ref 132b.
520 Chemical Reviews, 2009, Vol. 109, No. 2
Barbaro and Liguori
Scheme 8
Figure 3. Typical ESEM image of a DOWEX 50WX2 resin bead
containing the immobilized catalyst precursor [(-)-(TMBTP)Rh(NBD)]+ (TMBTP ) 4,4′-bis[diphenylphosphino]-2,2′,5,5′-tetramethyl-3,3′-bithiophene, NBD ) bicyclo[2.2.1]hepta-2,5-diene, backscattered electrons, 1200 magnifications, 20 KeV, 1 Torr). Reproduced
with permission from ref 59, page 5670. Copyright Wiley-VCH
Verlag GmbH & Co. KGaA.
RhCl(mtppms)3 immobilized on DEAE-Molselect resin
provided 1-phenylethanol from acetophenone in 15% yield
(60 °C, 80 bar H2, TOF ) 11.2 h-1, Scheme 7), which,
however, cannot compare with the activity observed in other
homogeneous catalyst systems.134a,137e
It is worth noticing that all reactions described in this
section, involving immobilization onto ion-exchange resin,
were carried out by packing the catalyst into a microfluidicsbased flow system (H-Cube), thus showing this heterogenization technology to be suitable for the engineering of
reactors and, eventually, for the scale-up to production
processes.138
Easy, in situ catalyst recycling is possible, featuring
negligible to very low rhodium leaching (0.3-2%), but with
the common trend of a drop in the catalytic activity in the
second and following cycles.
An in-depth investigation based on ESEM (environmental
scanning electron microscopy), EDS (energy dispersive X-ray
spectrometry), NMR (nuclear magnetic resonance), and GFAAS (graphite furnace atomic absorption spectroscopy)
analysis suggested that catalyst deactivation might be attributable to a binding interaction between the excess of
noninnocent sulfonato groups from the resin and the rhodium
centers after consumption of the substrate in the first
hydrogenation step (Rh coverage of anionic sites ca. 3%).149,150
However, the factors responsible for the activity decrease
in these systems have not been completely clarified to date.
Surface EDS microanalysis maps recorded on sections of
the catalyst beads showed the rhodium to be evenly
distributed within the solid support (see Figure 3 for an
electron image of [(-)-(TMBTP)Rh(NBD)]+ catalyst on
strong cation exchange resin). This evidence indicated that
the methanol, commonly used both for anchoring and
catalysis due to effective swelling of the resin in this solvent,
diffuses freely through the support, thus ensuring proper
access of the reactants to the active sites. The metal loss
was ascribed to catalyst decomposition due to fortuitous
ligand oxidation on the basis of 31P{1H} NMR experiments.
The above hypotheses were recently partially confirmed
by an asymmetric hydrogenation study carried out on methyl2-acetamidoacrylate using [((R)-Monophos)Rh(COD)]+ im-
3.3.1. Asymmetric Hydrogenations
Asymmetric hydrogenation is likely the most investigated
application of ion-exchange resin-supported molecular
catalysts.139-145 A survey of the literature on the topic was
released in 2005.59 Chiral homogeneous catalysts producing
optically active chemicals, with exceptionally high enantioselectivities and yields, are well established compounds
whose properties were widely described in the literature.146-148
Typically, they consist of noble metal complexes bearing
expensive, highly elaborated, single-enantiomer ligands.
Their recovery and recycle is, therefore, a major economical
need. The immobilization of these catalysts onto ionexchange resins is usually attained efficiently under smooth
conditions (metal loadings 0.3-1.5% w/w, using strong
cation exchange resins).
The heterogenized rhodium catalysts so obtained show
enantioselectivities comparable with those of the analogous
homogeneous phase reactions in the asymmetric hydrogenation of prochiral activated olefins, though with slower
reaction rates (Scheme 8). Representative literature examples
are summarized in Table 2.
Table 2. Asymmetric Hydrogenations Using Cationic Rh(I)-Chiral Diphosphine Ligand Catalysts Immobilized onto Ion-Exchange
Resinsa
heterogeneous
chiral ligand
R/R′
Ph-β-glup
Me/Ph
Propraphos
BDPP-(pNMe2)4
DIOP
TMBTP
Me/H
Me/Ph
Ph/Ph
Me/H
Me/H
a
-1
resin
TOF(h )
+
G-4-H
G-2-H+
G-1-H+
G-0.5-H+
Wofatit KP2
G-0.5-Li+
Nafion-NR-50
DOWEX 50WX2
DOWEX 50WX2
b
500
135b
60b
38b
15b
10b
400
38
52
ee (%)
94.1
94.9
95.5
95.3
94.3
84.6
50
54.6
99.9
homogeneous
TOF
(h-1)
b
ee (%)
6
91.1
0.9b
35b
1200
40
50
90.9
85.5
57
57.3
99.9
In methanol, room temperature, H2 pressure 1-5 bar. Substrate/catalyst ratio 100/1. b Half-life time t/2 (min).
ref
139
139
139
139
141
139
142
149
149
Ion Exchange Resins
Scheme 9
mobilized on Nafion (Scheme 9).151 There, the activity and
leaching of the supported catalyst were directly correlated
to the swelling of Nafion and to the solubility of the
molecular catalyst in the solvent used, respectively, by
comparison with other insoluble anionic supports: the former
was highest for water and methanol, while leaching was
lowest for water. Phosphotungstic acid on alumina was
shown to be the best support in this application, in terms of
activity and leaching, whereas Nafion was the poorest.
Negative effects due to immobilization by encapsulation
rather than on ionic interactions in Nafion were hypothesized
on these basis. Recycling was once again demonstrated with
almost complete retention of selectivity and activity over four
consecutive runs (water, 5 bar H2, 20 °C, conversion 100%,
TOF 17 h-1, ee 90%, Rh loss 0.01 mg/L. Homogeneous:
TOF 1700 h-1, ee 97%, CH2Cl2. Data from ref 151).
3.4. Oxidations
Previous examples of oxidations of organic substrates by
ion-exchange-supported homogeneous catalysts include the
oxidation of dihydrogen;152 the oxidation of hydrocarbons,153,154
alcohols, ketones and thiols;155-161 the hydroxylation of
aromatics;162,163 and the epoxidation of olefins, particularly
by immobilized metalloporphyrins.164-171 More recently,
applications in the field concentrated on the epoxidation and
dihydroxylation of alkenes, on photooxidation reactions, and
on other few examples of miscellaneous oxidations. These
are briefly discussed in the following sections.
3.4.1. Alkene Epoxidations
Dicationic Mn(III)-salen complexes immobilized on the
macroporous strong cation exchange resins Dowex MSC-1
(exchange capacity 4.5 mequiv/g) and Amberlite IRA-200
were used as heterogeneous epoxidation catalysts of different
alkenes by sodium periodate (Scheme 10).172,173
The ionic bonding between salen and the resins was so
strong that neither common organic solvents nor acidic or
electrolyte water solutions eluted the complex from the
support. Catalysts with metal loading of 0.063 mmol/g (1.4%
of exchange sites) proved to be quite efficient (TOF up to
50 h-1) and selective (up to 100%) in the epoxidation of
cyclic and linear alkenes, eventually bearing aromatic
substituents, in aqueous acetonitrile (Table 3). Analogously
to the parent homogeneous catalyst, a positive effect of a
nitrogen base axial coligand was observed. Catalysts’ reuse
was achieved with ease, though with a significant loss of
activity and leaching (<1%) in each run, which was ascribed
to the degradation of the manganese complex.
Chemical Reviews, 2009, Vol. 109, No. 2 521
Scheme 10
Questions may be raised on how the charge of a ionic
reagent (e.g., periodate) influences its penetration into the
interior of an ion-exchange resin. Indeed, at low ionic
strength and neutral pH values, anion transport through
Nafion membranes is known to be hindered by electrostatic
repulsion forces between the anions and the sulfonate fixedcharge sites in the membrane.174a However, lowering the pH
or increasing the ionic strength of a solution in contact with
strong cation exchange membranes causes the suppression
of electrostatic repulsions and a ready diffusion of anions
through the membrane.174b,c Diffusion coefficients through
Nafion were measured for Cl- and I- in the presence of their
zinc salts,174d for bisulfate anion in 0.001-1 M sulfuric
acid,174a and for bisulfate and Cr(VI) anions at high salt
concentrations (0.05-1.3 M) and low pH values.174e Diffusivities comparable with those observed in large pore alumina
ceramic diaphragms, with no negative fixed-charges, were
obtained in this latter case (ca. 0.8-1.1 × 10-6 cm2/s).
In the case at hand, i.e. periodate and fixed-charge
sulfonate resin, we must assume that the ionic strength
(NaIO4 0.13 M) is responsible for the smooth penetration of
the anionic oxidant into the resin and, thus, for the remarkable
activity observed. Replacement of the charged reactant with
uncharged ones (H2O2, tert-BuOOH) led invariably to lower
reaction rates, whereas addition of strongly basic amine
cocatalysts did not significantly affect the catalyst’s activity
(rate enhancement was observed for strongly coordinating
bases only).
The catalytic asymmetric epoxidation of unfunctionalized
olefins by meta-chloroperbenzoic acid was accomplished
using immobilized, sulfonato chiral salen Mn complex
(Scheme 11).175 A strong anion exchanger obtained from a
commercial Merrifield resin after quaternization with triTable 3. Examples of Alkene Epoxidations Using Mn(salen) on
Dowex MSC-1 Based Catalysts172
alkene
TOF (h-1)
selectivity (%)
cyclooctene
R-limonene
trans-stilbene
cis-stilbene
1-heptene
50
50
14
15
21
65 (1,2/8,9)
100 (trans)
55 (cis/trans)
522 Chemical Reviews, 2009, Vol. 109, No. 2
Barbaro and Liguori
Scheme 11
Scheme 13
Scheme 14
3.4.2. Asymmetric Dihydroxylations
Scheme 12
ethylamine was used as solid support (chloromethylated
styrene-divinylbenzene copolymer, 2% cross-linkage, gel
type).
The chiral epoxides were obtained faster, with higher
enantioselectivities, lower catalyst deactivation, and metal
leaching, compared both to the homogeneous catalyst and
to other noncovalently supported heterogeneous systems
using diverse porous materials (silica, Mg-Al-LDH). Selected
figures are summarized in Table 4. No explanations for the
observed behavior were given, however.
The resin-bound catalyst could be recovered by filtration
and reused without loss in ee and only minor changes in
activity. Data for 6-cyanochromene are reported in Scheme
12, as a representative example. Metal leaching was negligible in each run.
Table 4. Asymmetric Epoxidations of Olefins by m-CPBA Using
Immobilized Chiral salen Mn Catalysts175
alkene
support
styrene
silica
LDH
resin
trans-stilbene
silica
LDH
resin
TOF (h-1)
ee (%)
590
590
1250
1250
100
50
1250
1250
25
40
44
48
18
42
29
35
Recent examples of dihydroxylation of alkenes by exchangeresin-supported catalysts are limited to the asymmetric
version of this reaction (Scheme 13).176
Choudary et al. described an innovative system based on
osmate anions immobilized by ion-exchange on a strongbase exchanger. The resin used was the quaternary ammonium Merrifield polymer mentioned in the previous
section (Scheme 14).177,178 The final polymeric material
showed a metal content of 0.64 mmol Os/g and a BET
surface area of 317 m2/g. Reuse of the catalyst was further
motivated by the cost and toxicity of osmium in this case.
The heterogenized complex resin-OsO4 was used in the
osmium-assisted asymmetric dihydroxylation of olefins, using
various co-oxidants, to give chiral vicinal diols after in situ
complexation of the metal center with the chiral ligand
(DHQD)2PHAL ((DHQD)2PHAL ) dihydroquinidine 1,4phthalazinediyl diether) (Figure 4).
The process involves the known Os(VIII)/Os(VI) redox
cycle, in which the oxidation of resin-OsO4 by the co-oxidant,
followed by the interaction of the Os(VIII) oxo-complex with
the olefin and the ligand, forms the osmium(VI) monoglycolate ester, whose hydrolysis provides the diol, with
regeneration of the catalyst precursor.179 It is interesting to
note that neither OsO42- nor Os(VIII) species are leached
from the support during the catalytic reaction. This means
that Os(VI) is held on the resin by strong interactions and
that the Os(VIII) lifetime is too short to detach neutral OsO4
from the support. The possibility that Os(VIII) is firmly
anchored on the resin is ruled out by the fact that osmium
leached into solution when the reaction was carried out in
the absence of olefin.
Irrespective of the co-oxidant used (either N-methylmorpholine N-oxide (NMO), Fe(CN)63-, or O2), the ionexchanged catalyst showed superior efficiency, in terms of
both activity and selectivity, compared to analogous silica
and polymer bound recyclable systems (e.g., Kobayashi’s
ABS- and PEM-MC OsO4 catalysts).180 Data for the asymmetric dihydroxylation of methylstyrene are reported in Table
5 as a representative example.
This novel, favorable behavior was ascribed to proper
swelling of the resin in the reaction medium (H2O/t-BuOH),
Ion Exchange Resins
Chemical Reviews, 2009, Vol. 109, No. 2 523
Scheme 15
Scheme 16
Figure 4. Proposed mechanism for the resin-OsO4 catalyzed
asymmetric dihydroxylation of olefins. Co-oxidant, NMO; solvent,
t-BuOH/H2O; rt; chiral ligand ) (DHQD)2PHAL. See ref 177.
which enables easy access of all reactants to the metal active
sites inside the pores. Excellent yields and ee’s were observed
for acyclic, cyclic, and mono-, di-, and trisubstituted olefins.
The catalyst could be quantitatively recovered and reused
by simple filtration, showing consistently good and constant
efficiency over five runs, independently from the nature of
the co-oxidant and the solvent. Data for the asymmetric
dihydroxylation of R-methylstyrene are reported in Scheme
15, as an example.
3.4.3. Various Oxidations
Cytochrome P450 biomimetic model compounds based on
iron(III) porphyrins showed interesting nitric oxide synthaselike activity, i.e. the catalytic oxidation of L-arginine to nitric
oxide and L-citrulline by hydrogen peroxide (Scheme 16),
when supported on commercial ion exchange resins.181 The
strong cation exchanger Dowex 50WX8 (Na+ form, gel type)
was used to immobilize the cationic porphyrins FeTPAP
(meso-tetrakis(N,N,N,-trimethylammoniumphenyl)porphyrin iron(III) chloride) and FeTMPyP (meso-tetrakis(1-methylpyridinium-4-yl)porphyrin iron(III) chloride). The anionic
porphyrins FeTPPS (trisodiumdodecaaquatetrakis(p-sulfonatophenyl)porphyrinato iron(III)) and FeTCP (trisodium
bisaquatetrakis(p-carboxyphenyl)porphyrinato iron(III)) were
immobilized on the anion exchanger Dowex 1×8 (chloride
form, gel type). Heterogenization of metalloporphyrin cataTable 5. Asymmetric, Os-Supported Catalytic Dihydroxylations
of r-Methylstyrenea
support
TOF (h-1)
ee (%)
ref
b
7.8
7.4
7.5
3.4
4.1
91
90
90
76
78
177
177
177
180a
180b
resin
LDHb
silica
PEM-MCc
ABS-MCb
a
Chiral ligand (DHQD)2PHAL. b Co-oxidant NMO. c Co-oxidant
K3Fe(CN)6.
lysts was skillfully used to prevent catalyst deactivation by
µ-oxo dimers formation.182
Indeed, compared to the corresponding homogeneous
catalysts, the supported iron porphyrins are characterized by
both higher activity and stability. Yields 30 to 50 times higher
than those observed in the homogeneous phase were observed
with negligible metal leaching (typical TOF 68-9 and 1-0.3
h-1, for heterogeneous and homogeneous reactions, respectively). Contrarily, the soluble iron porphyrins were almost
completely oxidized by H2O2 under the same reaction
conditions (water, pH 7.8, 25 °C).
The photocatalyzed oxidation of alkanes and alkenes
was recently achieved through resin-supported tungsten
and platinum derivatives. Expensive platinum(II) bipyridine complexes loaded on IRA-200 exchanger were easily
recovered by simple filtration and recycled in the O2
oxidation of olefins, upon irradiation with visible light
(Scheme 17).183 The activity of the catalyst was very high,
providing [2 + 2], [4 + 2] cycloaddition or ene reaction
products, which were converted to the corresponding aldehydes, ester, or diols, depending on the substrate, in good
yields. The oxidation rates were dependent on the Pt loading
but decreased only slightly over ten consecutive runs.
Similarly, the known photooxidation catalyst decatungstate
W10O324-, whose immobilization on silica and polymeric
membranes was already explored,184-186 was anchored onto
anion exchange resins and used in the photocatalytic partial
oxidation of cyclohexane with molecular oxygen (Scheme
18).187
Cross-linked poly(4-vinylpyridine) methyl chloride quaternary salts, poly(4-vinylpyridinium tribromide) and
poly(4-vinylpyridinium p-toluenesulfonate), were used as
solid support (Scheme 19). The most effective catalyst
was that immobilized onto the cross-linked polymer (TON
) 4.41 at catalyst concentration 407 µmol dm-3, in
acetonitrile), with no detectable amount of metal released
into solution upon reuse, but with a decrease in the
turnover number of ca. 20% in two subsequent 6 h
reactions. Catalyst loading on the support affected the
524 Chemical Reviews, 2009, Vol. 109, No. 2
Scheme 17
Barbaro and Liguori
Scheme 20
3.5. Polymerizations
Scheme 18
Scheme 19
The development of supported catalysts for olefin polymerization and copolymerization is mainly motivated by the
cost and the environmental impact of products purification
using homogeneous catalysts and by the need to reuse the
expensive metal precursors, particularly in the scale-up to
production processes.189,190 A limited use of ion-exchange
resin supports was reported in recent years.
A hybrid nickel-copper catalyst was described consisting
of Ni2+ ions immobilized onto a macroporous, weak cation
exchange resin (cross-linked polyacrylate, carboxylate derivative, Ni loading 1.04 × 10-4 mol/ g resin) and a Cu(II)
complex (1 mol % CuCl2/tris[2-(dimethylamino)ethyl]amine), that was used in the controlled radical polymerization of methyl methacrylate.191 The catalyst showed high
activities, providing polymers with reasonable polydispersity
(Mw/Mn ) 1.26) at 90% conversion after 68 h (90 °C,
m-xylene, MMA/Ni ) 2000). The soluble copper cocatalyst
was added with the aim to achieve a better polymerization
control, due to its radical deactivator properties.192 The
heterogeneous nickel catalysts could be recovered by simple
centrifugation and recycled with essentially the same activity,
without any regeneration treatment. The metal residue in the
product polymer was quite low (Ni 7 ppm, Cu 29 ppm).
The synthesis of polyketones by copolymerization of
carbon monoxide and styrene was performed using a
sulfonated-resin-supported palladium catalyst, and its recyclability was explored.193 The immobilized catalyst showed
activities noticeably lower than that of the corresponding
homogeneous catalyst (143 g copolymer/g Pd-1 · h-1). Nonoptimized experiments proved that the catalyst can be reused,
though with a decrease in efficiency and with significant
metal leaching.
3.6. Miscellaneous
product selectivity but not the reaction rate. The lower
loading promoted cyclohexanone production whereas the
catalysts with high decatungstate content favored cyclohexyl hydroperoxide generation.
The degradative catalytic oxidation of 2,4-dichlorophenoxyacetic acid by H2O2 was accomplished with noticeable
efficiency (93% conversion after 30 min), in the presence
of ferrous ions immobilized on Amberlite IR-120 (sulfonic
acid exchanger, 8% cross-linked, gel type, 26 mg Fe/g resin),
UV light, and oxalate.188 The heterogenized catalyst could
be easily separated from the solution by filtration and
recycled. Enhancement of the reaction rate with reused
catalysts was attributed to massive Fe2+ leaching in solution
upon irradiation, however.
The selective, catalytic reduction of NO3- to N2 with
molecular hydrogen is an attractive way to purify low nitrate
content water to drinking limits.194 A bimetallic Pd-Cu
system based on metal(II) anionic chloro complexes immobilized onto a strong anion exchanger (Dowex 1×4, gel
type, tetraalkylammonium chloride) was used for this purpose
and characterized by several techniques, including SEM,
XRMA (X-ray microprobe analysis), XPRD (X-ray powder
diffraction), and TEM (transmission electron microscopy).195
The reduction of the immobilized complexes with NaBH4
in the liquid phase afforded the final catalyst in the form of
very small or amorphous metallic palladium and copper
nanodomains, evenly distributed within the resin beads
(loadings Pd 4.10% w/w, Cu 0.95% w/w, Scheme 20). These
were shown to be quite efficient (61% conversion after 135
min, 25 °C, water pH ) 6) and selective (N2 93.8%) in the
reduction of nitrates, albeit less active than the corresponding
alumina supported catalyst. Copper and palladium leaching
in solution was below the detection limit.
The immobilized chiral salen Mn complex described
previously in section 3.4.1 (Scheme 11) was successfully
Ion Exchange Resins
Scheme 21
employed in the catalytic oxidative kinetic resolution of
racemic secondary alcohols in water, using PhI(OAc)2 as
oxidant (Scheme 21).196
Conversions and enantiomeric excesses of the unreacted
alcohols were only slightly lower than those obtained in the
homogeneous phase catalysis, being highest for 1,2,3,4tetrahydro-1-naphtol (TOF 134 h-1, ee 96%). The selectivity
factors were similar but modest in any case (ranging from
2.4 to 9.2). The heterogenized catalyst was filtered and reused
with almost constant efficiency and without Mn leak into
solution.
4. Conclusions
Previous and recent findings on the use of ion-exchange
resins as support for the heterogenization of homogeneous
chemical catalysts provide clear evidence of the effectiveness
of this technology.
In one word, compared to other more sophisticated
immobilization strategies and materials, this approach is
characterized by its simplicity: easy to use and to recycle
heterogeneous catalysts are readily obtained from low-cost,
commercially available materials. The major benefits of the
method include the following:
(1)versatilitysin terms of variety of supported catalysts and
reactions performed;
(2)efficiencysthe activities and selectivities observed are
comparable, and sometimes even better, with those of the
free catalysts;
(3)quantitative recoverability;
(4)good stabilitysmechanical, thermal, and chemical;
(5)modularitysa proper combination of the resin type,
supported catalysts, and immobilization strategy allows a
systematic design of new heterogeneous catalysts.
Some broad conclusions concerning the activity and the
selectivity of resin-supported catalysts may be gathered from
literature examples. As a general rule, the activity of resinsupported catalysts is primarily governed by the swelling of
the resin in the reaction solvent, that is, by its porosity at
the microscale level and by its cross-linkage at the molecular
level. Proper swelling of the resin ensures efficient mass
transport throughout the support and faster reaction kinetics.
Additionally, use of gel-type, low cross-linked resins usually
provides good dispersions of the active species in the organic
polymer, which result in an enhanced stability and performance of the supported catalysts. Site-isolation effects may
also play an important role, especially in those processes
(e.g oxidations) in which the formation of inactive dimers
limits the catalyst functionality in solution. The interaction
of the molecular catalyst with either the functional sites or
the bulky polymer architecture of the resin may produce a
favorable, confined nanoenvironment which is not found in
the homogeneous phase.35,58,197
The effect of the ionic support on catalyst selectivity is
evident but more difficult to predict. Hindered access of
particular reaction intermediates to the catalysts’ active site
or stabilization of an unusual catalysts geometry, either by
steric constraints or by electrostatic interaction, may be
responsible for the selection of specific reaction pathways,
Chemical Reviews, 2009, Vol. 109, No. 2 525
when high selectivities are observed. The anion influence
on the enantioselectivity in catalytic asymmetric reactions
was previously demonstrated in solution.198 Further studies
aimed at elucidating these relationships in ionically immobilized catalysts are awaited in the future.
Drawbacks of ion-exchange resin are occasional and
limited to catalyst deactivation and/or metal leaching.
Advances in the field should mainly address these critical
issues. The functional groups bound to the polymer backbone
seem to play a role in the catalytic activity decrease, at least
in asymmetric hydrogenation reactions, where the catalytic
efficiency drop was associated with the strong interaction
between the sulfonato groups on the polymer and the metal
center, after the first hydrogenation cycle. The effect of the
counterion on the homogeneous phase reaction rates was
highlighted in the past,198,199 where the hydrogenation rate
was shown to decrease across the series (Al[OC(CF3)3]4)> [B(C6F5)4]- > PF6- . BF4- > CF3SO3-. This suggests
that the introduction of “innocent”, noncoordinating exchanger groups on the resin could be a useful strategy to
successfully combat activity loss in resin-supported catalysts.
The recent preparation of borate polystyrene colloids and
fluoro-borate polyelectrolytes has to be considered with great
interest in this regard.200,201
In order to gain high activities by resin-supported catalysts,
further aspects of innovation are foreseen in relation to the
materials shape and size. While the importance of obtaining
precise information on the morphology, size, pore volume,
and solvent mobility was already stressed before,63,202 other
macroscopic physical forms of porous ion-exchange resins
may be of great relevance in order to overcome the diffusion
limitations found with conventional beads, i.e. fibers203-207
and (open-celled) monoliths.208-212 A number of fibrous
materials are available in the form of woven or nonwoven
textiles, and some of them could be functionalized into ionexchange fibers by grafting sulfonic or quaternary ammonium
groups. With respect to powder and pellets, fibers have
flexible structures and excellent mass transfer characteristics.
They may find applications in liquid-phase operation, for
their low diffusion resistance, and in three-phase operation
for their low-pressure drop. Fibers may be used in several
forms, and hence, they are expected to be easier and cheaper
to install and disassemble than monolith-based reactors.
Porous monolith-type ion exchangers, either anion or
cation exchangers, are three-dimensional open-cellular structures with 5-50-µm diameter pores and porous volumes of
8-10 mL/g, instead of 0.5-2 mL/g for usual resins.
Compared with conventional exchangers, they have multiple
advantages: they are easier to pack in a column, they have
much higher ion exchange rates (faster equilibration), smaller
ion exchange band lengths (higher ion selectivity), higher
permeability, and remarkably improved mass transfer properties. Also, as the exchanger groups can be quantitatively
introduced, ion-exchange capacity can be accurately tuned.
These unique properties justify the many applications of resin
monoliths, which include electrodeionization, production of
ultrapure water, chromatography, and use as adsorbents. Use
of these materials could be very promising as support for
noncovalent catalyst immobilization: positive effects on
activity and selectivity are presumable.
Although not exactly pertinent to the topic of the present
review, it is worth noticing that catalytic systems based on
supported palladium(0) nanoparticles obtained by the reduction of Pd2+ species immobilized onto ion-exchange resins,
526 Chemical Reviews, 2009, Vol. 109, No. 2
which already have a number of precedents in the literature,213 have now reached a high level of development thanks
to the use of monolithic reactors based on polymer/glass
composites.214
Finally, composite materials were shown to be helpful to
reduce catalyst leaching; fine-tuning of their properties may
contribute significantly to accomplish this goal.215-218
All the above considerations suggest that a new technology
based on ion-exchange resin is nearly mature for the
development of synthetic processes beyond the mere laboratory scale. A rational design of a new generation of (flow)
reactors based on these materials is possible, including
catalytic membrane reactors.219-221 Examples in this review
show that such immobilized catalysts are ideally suited for
assembling into reactors and are already in operation for
liquid-phase applications. Commercial processes based on
heterogenized homogeneous catalyst are still rare: it is
noteworthy that one of these, i.e., the Acetica process,
involves the use of ion-exchange resins.
5. Acknowledgments
Thanks are due to the IDECAT Network of Excellence,
to the FP6 European Community (Contract NMP3-CT-2005011730, www.idecat.org), and to the EBH2 project granted
by the POR Ob. 3 2000/2006, Regione Toscana, Italy.
6. References
(1) Anastas, P. T.; Warner, J. C. Green Chemistry: Theory and Practice;
Oxford University Press: Oxford, 1998.
(2) Horváth, I. T.; Anastas, P. T. Chem. ReV. 2007, 107, 2169.
(3) Anastas, P. T.; Kirchhoff, M. M.; Williamson, T. C. Appl. Catal.,
A: Gen. 2001, 221, 3.
(4) Sheldon, R. A. Green Chem. 2008, 10, 359.
(5) www.epa.gov/greenchemistry.
(6) (a) A European Technology Platform For Sustainable Chemistry,
http://www.cefic-sustech.org/files/Publications/ETP_sustainable_chemistry.pdf. (b) SusChem, Implementation Action Plan, 2006, http://
www.suschem.org.
(7) Applied Homogeneous Catalysis with Organometallic Compounds;
Cornils, B., Herrmann, W. A., Eds.; Wiley-VCH: Weinheim, 2002.
(8) Asymmetric Catalysis on Industrial Scale; Blaser, H. U., Schmidt,
E., Eds.; Wiley-VCH: Weinheim, 2004.
(9) Stinson, S. C. Chem. Eng. News 2001, (October 1), 79.
(10) Machado, R. M.; Heier, K. R.; Broekhuis, R. R. Curr. Opin. Drug
DiscoVery DeV. 2001, 4, 745.
(11) Johnson, N. B.; Lennon, I. C.; Moran, P. H.; Ramsden, J. A. Acc.
Chem. Res. 2007, 40, 1291.
(12) Klingler, F. D. Acc. Chem. Res. 2007, 40, 1367.
(13) Saudan, L. A. Acc. Chem. Res. 2007, 40, 1309.
(14) Briant D. R. Classical Homogeneous Catalyst Separation Technology.
In Catalyst Separation, RecoVery and Recycling; Chemistry and
Process Design; Cole-Hamilton, D. J., Tooze, R. P., Eds.; Springer:
Dordrecht, Chapter 2, 2006.
(15) Livingston, A.; Peeva, L.; Han, S.; Nair, D.; Luthra, S. S.; White,
L. S.; Freitas Dos, S.; Luisa, M. Ann. N. Y. Acad. Sci. 2003, 984,
123.
(16) Lu, J.; Lazzaroni, M. J.; Hallett, J. P.; Bommarius, A. S.; Liotta,
C. L.; Eckert, C. A. Ind. Eng. Chem. Res. 2004, 43, 1586.
(17) Cole-Hamilton, D. J.; Tooze, R. P. Homogeneous CatalysissAdVantages
and Problems, in Catalyst Separation, RecoVery and Recycling;
Chemistry and Process Design; Cole-Hamilton, D. J., Tooze, R. P.,
Eds.; Springer: Dordrecht, Chapter 1, 2006.
(18) Thomas, J. M.; Thomas, W. J. Principles and Practice of Heterogeneous Catalysis; Wiley-VCH: Weinheim, 1996.
(19) Fine Chemicals Through Heterogeneous Catalysis; Sheldon, R. A.,
van Bekkum, H., Eds.; Wiley-VCH: Weinheim, 2001.
(20) Blaser, H. U.; Pugin, B.; Studer, M. EnantioselectiVe Heterogeneous
Catalysis: Academic and Industrial Challenges in Chiral Catalyst
Immobilisation and Recycling; De Vos, D. E., Vankelecom, I. F. J.,
Jacobs, P. A., Eds.; Wiley-VCH: Weinheim, 2000; Chapter 1.
(21) Copéret, C.; Chabanas, M.; Saint-Arroman, R. P.; Basset, J. M.
Angew. Chem., Int. Ed. 2003, 42, 156.
Barbaro and Liguori
(22) Kirschning, A.; Jas, G. Applications of Immobilized Catalysts in
Continuous Flow Processes; Topics in Organometallic Chemistry
Vol. 242; Springer: Berlin/Heidelberg, 2004.
(23) End, N.; Schöning, K. U. Immobilized Catalysts in Industrial
Research and Application; Topics in Organometallic Chemistry Vol.
242; Springer: Berlin/Heidelberg, 2004.
(24) Corain, B.; Centomo, P.; Zecca, M. Chim. Ind. (Milano) 2004, 86,
114.
(25) Ding, K. Pure Appl. Chem. 2006, 78, 293.
(26) Gladysz, J. A. Chem. ReV. 2002, 102, 3215.
(27) Reek, J. N. H.; Van Leeuven, P. W. N. M.; Van Der Ham, A.G. J.;
De Haan, A. B. Supported Catalysts. In Catalyst Separation, RecoVery
and Recycling; Chemistry and Process Design,; Cole-Hamilton, D. J.,
Tooze, R. P., Eds.; Springer: Dordrecht, 2006, Chapter 3.
(28) Chiral Catalyst Immobilisation and Recycling; De Vos, D. E.,
Vankelecom, I. F. J., Jacobs, P. A., Eds.; Wiley-VCH: Weinheim,
2000.
(29) Homogeneous, Single-site Heterogeneous and Nanostructured Catalysts for Sustainable DeVelopment; Giambastiani, G., Bianchini, C.,
Eds.;Topics in Catalysis Vol. 40; Springer: Dordrecht, 2006.
(30) Bell, A. T. Science 2003, 299, 1688.
(31) Cole-Hamilton, D. J. Science 2003, 299, 1702.
(32) Baker, R. T.; Kobayashi, S.; Leitner, W. AdV. Synth. Catal. 2006,
348, 1337.
(33) Fan, Q. H.; Li, Y. M.; Chan, A. S. C. Chem. ReV. 2002, 102, 3385.
(34) Kirschning, A.; Solodenko, W.; Mennecke, K. Chem.sEur. J. 2006,
12, 5972.
(35) Song, C. E.; Kim, D. H.; Choi, D. S. Eur. J. Inorg. Chem. 2006,
2927.
(36) Bianchini, C.; Barbaro, P. Top. Catal. 2002, 19, 17.
(37) Waller, F. J. Chem. Ind. 2003, 89, 1.
(38) Müller, C.; Nijkamp, M. G.; Vogt, D. Eur. J. Inorg. Chem. 2005,
4011.
(39) Supported Metal Catalysts and Their Applications; Sherrington, D. C.,
Kybett, A. P., Eds.; Royal Society of Chemistry: Cambridge, 2000.
(40) Anderson, S.; Yang, H.; Tanielyan, S. K.; Augustine, R. L. Chem.
Ind. 2001, 82, 557.
(41) Song, C. E.; Lee, S. G. Chem. ReV. 2002, 102, 3495.
(42) De Vos, D. E.; Dams, M.; Sels, B. F.; Jacobs, P. A. Chem. ReV.
2002, 102, 3615.
(43) Van Heerbeek, R.; Kamer, P. C. J.; Van Leeuwen, P. W. N. M.;
Reek, J. N. H. Chem. ReV. 2002, 102, 3717.
(44) McNamara, C. A.; Dixon, M. J.; Bradley, M. Chem. ReV. 2002, 102,
3275.
(45) Benaglia, M.; Puglisi, A.; Cozzi, F. Chem. ReV. 2003, 103, 3401.
(46) Wight, A. P.; Davis, M. E. Chem. ReV. 2002, 102, 3589.
(47) Valkenberg, M. H.; Hölderich, W. F. Catal. ReV. 2002, 44, 321.
(48) Alaerts, L.; Wahlen, J.; Jacobs, P. A.; De Vos, D. E. Chem. Commun.
2008, 1727.
(49) Bein, T. Curr. Opin. Solid State Mater. Sci. 1999, 59, 73.
(50) Hodge, P. Chem. Soc. ReV. 1997, 26, 417.
(51) Augustine, R. L.; Tanielyan, S. K.; Mahata, N.; Gao, Y.; Zsigmond,
A.; Yang, H. Appl. Catal., A: Gen. 2003, 256, 69.
(52) Simons, C.; Hanefeld, U.; Arends, I. W. C. E.; Sheldon, R. A.;
Maschmeyer, T. Chem.sEur. J. 2004, 10, 5829.
(53) Jones, M. D.; Raja, R.; Thomas, J. M.; Johnson, B. F. G. Top. Catal.
2003, 25, 71.
(54) Thomas, J. M.; Raja, R. The AdVantages and future Potential of
Single-site Heterogeneous Catalysis in Homogeneous, Single-site
Heterogeneous and Nanostructured Catalysts for Sustainable DeVelopment; Giambastiani, G., Bianchini, C., Eds.;Topics in Catalysis
Vol. 40; Springer: Dordrecht, 2006.
(55) McMorn, P.; Hutchings, G. J. Chem. Soc. ReV. 2004, 33, 108.
(56) Fraile, J. M.; Garcı́a, J. I.; Mayoral, J. A. Non-coValent Immobilization
of Catalysts Based on Chiral Diazaligands; Topics in Organometallic
Chemistry Vol. 15; Springer: Berlin/Heidelberg, 2005.
(57) Horn, J.; Michalek, F.; Tzschucke, C. C.; Bannwarth, W. NonCoValently Solid-Phase Bound Catalysts for Organic Synthesis;
Topics in Organometallic Chemistry Vol. 242; Springer: Berlin/
Heidelberg, 2004.
(58) Single-site heterogeneous catalysts are commonly defined as “solid
catalysts where individual isolated atoms, ions, molecular complexes,
including asymmetric organometallic species, are firmly and uniformly anchored to a support”. See: Thomas, J. M.; Raja, R.; Lewis,
D. W. Angew. Chem., Int. Ed. 2005, 44, 6456.
(59) Barbaro, P. Chem.s Eur. J. 2006, 12, 5666.
(60) Vankelecom, I. F. J. Chem. ReV. 2002, 102, 3779.
(61) Chesney, A. Green Chem. 1999, 1, 209.
(62) Hölscher, M. Green Chem. 2006, 8, 761.
(63) Gelbard, G. Ind. Eng. Chem. Res. 2005, 44, 8468.
(64) Seen, A. J. J. Mol. Catal., A: Chem. 2001, 177, 105.
(65) Brulé, E.; De Miguel, Y. R. Org. Biomol. Chem. 2006, 4, 599.
Ion Exchange Resins
(66) Bickerstaff, G. F., Ed. Immobilization of Enzymes and Cells; Humana
Press: Totowa, NJ, 1997.
(67) Pedersen, S.; Christensen, M. W. Immobilized Biocatalysts. In
Applied Biocatalysis, 2nd ed.; Straathof, A. J. J., Adlercreutz, P.,
Eds.; Harwood Academic Publishers: Amsterdam, 2000; pp 213228.
(68) Harmer, M. A.; Sun, Q. Appl. Catal., A: Gen. 2001, 221, 45.
(69) Sharma, M. M. React. Funct. Polym. 1995, 26, 3.
(70) Tanabe, K.; Hölderich, W. F. Appl. Catal., A 1999, 181, 399.
(71) Harmer, M. A. Industrial Processes Using Solid Acid Catalysts. In
Handbook of Green Chemistry and Technology; Clark, J., Macquarrie,
D., Eds.; Blackwell Science: Oxford, 2002; pp 86-119.
(72) Hubicki, Z.; Leszczynska, M. Przem. Chem. 2005, 84, 750.
(73) Rangasamy, P.; Thomas, Z. S.; Kurt, S. E. U. S. Patent, US 5208194,
1993.
(74) Trens, P.; Peckett, J. W.; Stathopoulos, V. N.; Hudson, M. J.;
Pomonis, P. J. Appl. Catal., A: Gen. 2002, 241, 217.
(75) Trens, P.; Caps, V.; Peckett, J. W. Appl. Catal., A: Gen. 2003, 251,
19.
(76) (a) Zagorodni, A. A. Ion Exchange Materials: Properties and
Applications; Elsevier: Amsterdam, 2006. (b) Okay, O. Prog. Polym.
Sci. 2000, 25, 711.
(77) Ettre, L. S. Pure Appl. Chem. 1993, 65, 819.
(78) Marton, A. Pure Appl. Chem. 1997, 69, 1481.
(79) Streat, M. In AdVances in Ion Exchange for Industry and Research;
Williams, P. A., Dyer A., Eds.; The Royal Society of Chemistry:
Cornwell, U.K., 1998.
(80) Montgomery, J. M. Water Treatment Principals and Design; John
Wiley & Sons: New York, 1985.
(81) Sata, T. Pure Appl. Chem. 1986, 58, 1613.
(82) Nachod, F. C.; Schubert, J. Ion Exchange Technology; Academic
Press: New York, NY, 1956.
(83) Helfferich, F. Ion Exchange; McGraw-Hill Book Company: New
York, NY, 1962.
(84) Gierke, T. D.; Munn, G. E.; Wilson, F. C. J. Polym. Sci., Polym.
Phys. Ed. 1981, 19, 1687.
(85) Iwamoto, R.; Oguro, K.; Sato, M.; Iseki, Y. J. Phys. Chem. B 2002,
106, 6973.
(86) Fujimura, M.; Hashimoto, T.; Kawai, H. Macromolecules 1982, 15,
136.
(87) Seen, A. J. J. Mol. Catal., A: Chem. 2001, 177, 105.
(88) Waller, F. J. Catal. ReV. Sci. Eng. 1986, 28, 1.
(89) Ahmed, M.; Malik, M. A.; Pervez, S.; Raffiq, M. Eur. Polym. J.
2004, 4, 1609.
(90) Selke, R.; Häupke, K.; Krause, W. J. Mol. Catal. 1989, 56, 315.
(91) Samsonov, G. V.; Pasechnik, V. A. Russ. Chem. ReV. 1969, 38, 547.
(92) Bogatyrev, V. L.; Sokolova, N. P. Russ. Chem. Bull. 1969, 18, 1573.
(93) Haag, R.; Hegel, A.; Strumblé, J. F. In Handbook of Combinatorial
Chemistry; Hanko, R., Nicolau, K. C., Hartwig, W., Eds.; WileyInterscience: New York, 2002.
(94) (a) Jerome, S. M. Sci. Total EnViron. 1988, 70, 275. (b) Van der
Walt, T. N.; Strelow, F. W. E. Anal. Chem. 1983, 55, 212. (c) Siemer,
D. D. Anal. Chem. 1980, 52, 1874.
(95) Yu, L.; Chen, D.; Li, J.; Wang, P. G. J. Org. Chem. 1997, 62, 3575.
(96) (a) Kobayashi, S.; Nagayama, S. J. Org. Chem. 1996, 61, 2256. (b)
Kobayashi, S.; Nagayama, S. J. Am. Chem. Soc. 1996, 118, 8977.
(97) Dioos, B. M. L.; Vankelecom, I. F. J.; Jacos, P. A. AdV. Synth. Catal.
2006, 348, 1413.
(98) Corma, A. Catal. ReV. 2004, 46, 369.
(99) Simons, C.; Hanefeld, U.; Arends, I.W.C.E.; Minnaard, A. J.;
Maschmeyer, T.; Sheldon, R. A. Chem. Commun. 2004, 2830.
(100) European Agency for the Evaluation of Medicinal Products, Note
for guidance on specification limits for residues of metal catalysts,
CPMP/SWP/QWP/4446/00, June 2002.
(101) Cornils, B.; Herrmann, W. A. Aqueous-Phase Organometallic
Catalysis, 2nd ed.; Wiley-VCH: Weinheim, 2004.
(102) Joo,′ F. Aqueous Organometallic Catalysis; Kluwer:Dordrecht, 2001.
(103) Yoneda, N.; Kusano, S.; Yasui, M.; Pujado, P.; Wilcher, S. Appl.
Catal., A: Gen. 2001, 221, 253.
(104) Howard, M. J.; Jones, M. D.; Roberts, M. S.; Taylor, S. A. Catal.
Today 1993, 18, 325.
(105) (a) Dekleva, T. W.; Foster, D. AdV. Catal. 1986, 34, 81. (b) Foster,
D. AdV. Organomet. Chem. 1979, 17, 255. (c) Maitlis, P. M.; Haynes,
A.; Sunley, G. J.; Howard, M. J. J. Chem. Soc., Dalton Trans. 1996,
2187. (d) Haynes, A.; Mann, B. E.; Morris, G. E.; Maitlis, P. M.
J. Am. Chem. Soc. 1993, 115, 4093.
(106) Howard, M. J.; Jones, M. D.; Roberts, M. S.; Taylor, S. A. Catal.
Today 1993, 18, 325.
(107) (a) Jiang, D. Z.; Li, X. B.; Wang, E. L. Macromol. Symp. 1996, 105,
161. (b) Sowden, R. J.; Sellin, M. F.; De Blasio, N.; Cole-Hamilton,
D. J. Chem. Commun. 1999, 2511. (c) De Blasio, N.; Wright, M. R.;
Mazzocchia, C.; Cole-Hamilton, D. J. J. Organomet. Chem. 1998,
551, 229. (d) De Blasio, N.; Kaddouri, A.; Mazzocchia, C.; Cole-
Chemical Reviews, 2009, Vol. 109, No. 2 527
(108)
(109)
(110)
(111)
(112)
(113)
(114)
(115)
(116)
(117)
(118)
(119)
(120)
(121)
(122)
(123)
(124)
(125)
(126)
(127)
(128)
(129)
(130)
(131)
(132)
(133)
(134)
Hamilton, D. J. J. Catal. 1998, 176, 253. (e) MarstonC. R.; GoeG.
L. E.P. 0277824, 1988. (f) Scates, M. O.; Warner, R. J.; Torrence,
G. P. E.P. 0656811. 1995.
Drago, R. S.; Nyberg, E. D.; El A’mma, A.; Zombeck, A. Inorg.
Chem. 1981, 20, 641.
Drago, R. S.; El A’mma, A. U.S. Patent 4,328,125, 1982.
Yoneda, N.; Shiroto, Y.; Hamato, K.; Asaoka, S.; Maejima, T. U. S.
Patent 5,334,755, 1994.
Yoneda, N.; Minami, T.; Weiszmann, J.; Spehlmann, B. In The
Chiyoda/UOP Acetica TM Process: A NoVel Acetic Acid Technology
in Studies; Hattori, H., Otsuka, K., Eds.; Surface Science and Catalysis
Vol. 121; Kodansha: Tokyo, 1999; pp 93-98.
Yoneda, N.; Minami, T.; Hamato, K.; Shiroto, Y.; Hosono, Y. J.
Jpn. Petrol. Inst. 2003, 46, 229.
Malik, M. A.; Naheed, R. e-Polym. 2007, 135, 1.
Yoneda, N.; Hosono, Y. J. Chem. Eng. Jpn. 2004, 37, 536.
Haynes, A.; Maitlis, P. M.; Quyoum, R.; Pulling, C.; Adams, H.;
Spey, S. E.; Strange, R. W. J. Chem. Soc., Dalton Trans. 2002, 2565.
Giammatteo, M.; Tauro, L.; D’Archivio, A. A.; Galantini, L.; Panatta,
A.; Tettamanti, E.; Jerabek, K.; Corain, B. J. Mol. Catal., A: Chem.
2007, 268, 176.
http://www.chiyoda-corp.com/biz/e/hpi/acetica.shtml.
Futamura, C.; Hosono, Y.; Minami, T. Chem. Eng. Tokyo 2004, 49,
127.
Yoneda, N.; Shiroto, Y.; Hamato, K.; Asaoka, S.; Maejima, T. E.P.
0567331, 1993.
(a) Rhodium Catalyzed Hydroformylation; van Leeuwen, P. W. N. M.;
Claver, C., Eds.; Springer: Dordrecht, 2007. (b) Evans, D.; Osborn,
J. A.; Wilkinson, G. J. Chem. Soc. A 1968, 3133.
(a) Hintermair, U.; Zhao, G. Y.; Santini, C. C.; Muldoon, M. J.; ColeHamilton, D. J. Chem. Commun. 2007, 1462. (b) Pagar, N. S.;
Deshpande, R. M.; Chaudhari, R. V. Catal. Lett. 2006, 110, 129.
(a) Bryant, D. E.; Kilner, M. J. Mol. Catal., A: Chem. 2003, 193,
83. (b) Balué, J.; Bayon, J. C. J. Mol. Catal., A: Chem. 1999, 137,
193. (c) Toth, I.; Hanson, B. E.; Guo, I.; Davis, M. Catal. Lett. 1991,
8, 209.
(a) Kuntz, E. G. CHEMTECH 1987, 17, 570. (b) Cornils, B.; Kuntz,
E. J. Organomet. Chem. 1995, 502, 177. (c) Dyson, P. J.; Ellis, D. J.;
Welton, T. Platinum Met. ReV. 1998, 42, 135. (d) Huang, Y.; Min,
L.; Li, Y.; Li, R.; Cheng, P.; Li, X. Catal. Commun. 2002, 3, 71.
Diwakar, M. M.; Deshpande, R. M.; Chaudhari, R. V. J. Mol. Catal.,
A: Chem. 2005, 232, 179.
Westerterp, K. R.; Molga, E. J.; Van Gelder, K. B. Chem. Eng. Proc.
1997, 36, 17.
Nishimura, S. Handbook of Heterogeneous Catalytic Hydrogenation
for Organic Synthesis; Wiley-VCH: Weinheim, 2001.
Shimizu, H.; Nagasaki, I.; Matsumura, K.; Sayo, N.; Saito, T. Acc.
Chem. Res. 2007, 40, 1385.
Cornils, B.; Wiebus, E. CHEMTECH 1995, 25, 33.
Kohlpaintner, C. W.; Fischer, R. W.; Cornils, B. Appl. Catal., A:
Gen. 2001, 221, 219.
Joó, F.; Kovátcs, J.; Kathó, Á.; Bényei, A. C.; Decuir, T.; Darensbourg, D. J. Inorg. Synth. 1998, 32, 1.
Horváth, H. H.; Papp, G.; Csajági, C.; Joó, F. Catal. Commun. 2007,
8, 442.
(a) Horváth, H. H.; Joó, F. React. Kinet. Catal. Lett. 2005, 85, 355.
(b) Sánchez-Delgado, R.; Medina, M.; López-Linares, F.; Fuentes,
A. J. Mol. Catal., A: Chem. 1997, 116, 167. (c) Joó, F.; Kovács, J.;
Bényei, A. Cs.; Kathó, Á. Angew. Chem., Int. Ed. 1998, 37, 969. (d)
Andriollo, A.; Carrasquel, J.; Mariño, J.; López, F. A.; Páez, D. E.;
Rojas, I.; Valencia, N. J. Mol. Catal., A: Chem. 1997, 116, 157. (e)
Joó, F.; Kovács, J.; Bényei, A. Cs.; Kathó, Á. Catal. Today 1998,
42, 441. (f) Grosselin, J. M.; Mercier, C.; Allmang, G.; Grass, F.
Organometallics 1991, 10, 2126. (g) Wright, G.; Bjerrum, J. Acta
Chem. Scand. 1962, 16, 1261. (h) Evans, D.; Osborn, J. A.; Jardine,
F. H.; Wilkinson, G. Nature 1965, 208, 1203.
(a) Ojima, I.; Eguchi, M.; Tzamarioudaki, M. In ComprehensiVe
Organometallic Chemistry II; Abel, E. W., Stone, F. G. A.,
Wilkinson, G., Eds.; Vol. 12 Transition Metal Organometallics in
Organic Synthesis; Pergamon: Oxford, 1995. (b) Noyori, R.; Ohkuma,
T.; Kitamura, M.; Takaya, H.; Sayo, N.; Kumobayashi, H.; Akutagawa, S. J. Am. Chem. Soc. 1987, 109, 5856. (c) James, B. R.
Homogeneous Hydrogenation; Wiley: New York, 1973. (d) Noyori,
R.; Hashiguchi, S. Acc. Chem. Res. 1997, 30, 97. (e) Ohkuma, T.;
NoyoriR. In Transition Metals for Organic Synthesis: Building Blocks
and Fine Chemicals, Vol. 2; Beller, M., BolmC., Eds.; Wiley-VCH:
Weinheim, 1998.
(a) Noyori, R.; Ohkuma, T. Angew. Chem., Int. Ed. 2001, 40, 41.
(b) Augustine, R. L. AdV. Catal. 1976, 25, 56. (c) Keinan, E.;
GreenspoonN. In ComprehensiVe Organic Synthesis, Vol. 8; Trost,
B. M., FlemingI., Eds.; Pergamon: Oxford, 1991. (d) Chaloner, P. A.;
Esteruelas, M. A.; Joó, F.; Oro, L. A. Homogeneous Hydrogenation;
Kluwer: Dordrecht, 1994.
528 Chemical Reviews, 2009, Vol. 109, No. 2
(135) Note: A similar biphasic system based on a TPPTS-RuCl3 catalyst
provided cinnamyl alcohol with a selectivity of 98% and a conversion
greater that 95%. See ref 132f.
(136) (a) Tani, K.; Tanigawa, E.; Tatsuno, Y.; Otsuka, S. J. Organomet.
Chem. 1985, 279, 87. (b) Huang, Y. H.; Gladysz, J. A. J. Chem.
Educ. 1988, 65, 298. (c) Bryndza, H. E.; Tam, W. Chem. ReV. 1988,
88, 1163.
(137) (a) Ikariya, T.; Murata, K.; Noyori, R. Org. Biomol. Chem. 2006, 4,
393. (b) Sandoval, C. A.; Ohkuma, T.; Muñiz, K.; Noyori, R. J. Am.
Chem. Soc. 2003, 125, 13490. (c) Ohkuma, T.; Ooka, H.; Hashiguchi,
S.; Ikariya, T.; Noyori, R. J. Am. Chem. Soc. 1995, 117, 2675. (d)
Ohkuma, T.; Ooka, H.; Ikariya, T.; Noyori, R. J. Am. Chem. Soc.
1995, 117, 10417. (e) Noyori, R.; Ohkuma, T. Pure Appl. Chem.
1999, 71, 1493.
(138) http://www.thalesnano.com/files/file/brochures/applications/ThalesNano_ Catalyst _Screening.pdf.
(139) (a) Selke, R.; Capka, M. J. Mol. Catal. 1990, 63, 319. (b) Selke, R.;
Häupke, K.; Krause, W. J. Mol. Catal. 1989, 56, 315.
(140) Selke, R.; Pracejus, H. J. Mol. Catal. 1986, 37, 213.
(141) Selke, R. J. Mol. Catal. 1986, 37, 227.
(142) Tóth, I.; Hanson, B. E. J. Mol. Catal. 1992, 71, 365.
(143) Tóth, I.; Hanson, B. E.; Davis, M. E. J. Organomet. Chem. 1990,
396, 363.
(144) Tóth, I.; Hanson, B. E.; Davis, M. E. J. Organomet. Chem. 1990,
397, 109.
(145) Brunner, H.; Bielmeier, E.; Wiehl, J. J. Organomet. Chem. 1990,
384, 223.
(146) Catalytic Asymmetric Synthesis, 2nd ed.; Ojima, I., Ed.; Wiley-VCH:
Weinheim, 2000.
(147) Noyori, R. Asymmetric Catalysis in Organic Synthesis; John Wiley
& Sons: New York, 1994.
(148) Tang, W.; Zhang, X. Chem. ReV. 2003, 103, 3029.
(149) Barbaro, P.; Bianchini, C.; Giambastiani, G.; Oberhauser, W.; Morassi
Bonzi, L.; Rossi, F.; Dal Santo, V. J. Chem. Soc., Dalton Trans.
2004, 1783.
(150) Evans, S. Chem. Sci. 2004, 1, C51.
(151) Simons, C.; Hanefeld, U.; Arends, I. W. C. E.; Maschmeyer, T.;
Sheldon, R. A. J. Catal. 2006, 239, 212.
(152) Blanco-Brieva, G.; Cano-Serrano, E.; Campos-Martin, J. M.; Fierro,
J. L. G. Chem. Commun. 2004, 1184.
(153) Yatsimirskii, A.; Erokhin, A. S.; Berezin, I. V. IzV. Akad. Nauk SSSR,
Ser. Khim. 1980, 42.
(154) Zhenghong, G.; Jing, S. Y.; Siping, W.; Ping, Y. Appl. Catal., A
2001, 209, 27.
(155) Kanemoto, S.; Saimoto, H.; Oshima, K.; Nozaki, H. Tetrahedron
Lett. 1984, 25, 3317.
(156) Kanemoto, S.; Saimoto, H.; Oshima, K.; Utimoto, K.; Nozaki, H.
Bull. Chem. Soc. Jpn. 1989, 62, 519.
(157) Sun, D.; Zhong, J.; Chen, H.; Chen, J. Hubei Daxue Xuebao, Ziran
Kexueban 1999, 21, 158 (CAN 132: 92817).
(158) Skorobogaty, A.; Smith, T. D. J. Mol. Catal. 1982, 16, 131.
(159) Mifune, M.; Harada, R.; Iwado, A.; Motohashi, N.; Saito, Y. Talanta
1998, 46, 1583.
(160) Iwado, A.; Mifune, M.; Harada, R.; Mukuno, T.; Motohashi, N.; Saito,
Y. Anal. Sci. 1998, 14, 515.
(161) Palazzi, C.; Pinna, F.; Strukul, G. J. Mol. Catal., A: Chem. 2000,
151, 245.
(162) Chen, J.; Cao, S.; Tan, B. Lizi Jiaohuan Yu Xifu 1999, 15, 263 (CAN
132:124439).
(163) Yu, J.; Liu, Q.; Zhang, W.; Wenxiang, J.; Jiang, Y.; Wu, T.; Sun, J.,
(CAN 135:92403).
(164) Gelbard, G.; Breton, F.; Sherrington, D. C.; Quenard, M. J. Mol.
Catal. 2000, 153, 7.
(165) Kotov, S.; Boeva, R.; Yordanov, N. Oxid. Commun. 1984, 6, 55 (CAN
101:229740).
(166) Kotov, S.; Boeva, R.; Yordanov, N. J. Mol. Catal., A 1999, 139,
271.
(167) Leanord, D. R.; Lindsay Smith, J. R. J. Chem. Soc., Perkin Trans. 2
1991, 25.
(168) Iwado, A.; Mifune, M.; Kato, J.; Oda, J.; Chikuma, M.; Motohashi,
N.; Saito, Y. Chem. Pharm. Bull. 2000, 48, 1831.
(169) Campestrini, S.; Meunier, B. Inorg. Chem. 1992, 31, 1999.
(170) Sacco, H. C.; Iamamoto, Y.; Lindsay Smith, J. R. J. Chem. Soc.,
Perkin Trans. 2 2001, 181.
(171) Idemitsu Kosan Co. Jpn. Patent JP 56,081,137, 1981.
(172) Bahramian, B.; Mirkhani, V.; Moghadam, M.; Tangestaninejad, S.
Appl. Catal., A: Gen. 2006, 301, 169.
(173) Mirkhani, V.; Moghadam, M.; Tangestaninejad, S.; Bahramian, B.
Monatsh. Chem. 2007, 138, 1303.
(174) (a) Verbrugge, M. W.; Hill, R. F. J. Electrochem. Soc. 1990, 137,
893. (b) Capeci, S. W.; Pintauro, P. N.; Bennion, D. N. J.
Electrochem. Soc. 1989, 136, 2876. (c) Pourcelly, G.; Lindheimer,
A.; Gavach, C. J. Electroanal. Chem. 1991, 305, 97. (d) Sodaye,
Barbaro and Liguori
(175)
(176)
(177)
(178)
(179)
(180)
(181)
(182)
(183)
(184)
(185)
(186)
(187)
(188)
(189)
(190)
(191)
(192)
(193)
(194)
(195)
(196)
(197)
(198)
(199)
(200)
(201)
(202)
(203)
(204)
(205)
(206)
(207)
H. S.; Pujari, P. K.; Goswami, A.; Manohar, S. B. J. Radioanal.
Nucl. Chem. 1996, 214, 399. (e) Huang, K. L.; Holsen, T. M.; Selman,
J. R. Ind. Eng. Chem. Res. 2003, 42, 3620.
Choudary, B. M.; Ramani, T.; Maheswaran, H.; Prashant, L.;
Ranganath, K. V. S.; Vijay Kumar, K. AdV. Synth. Catal. 2006, 348,
493.
(a) Kolb, H. C.; Van Nieuwenhze, M. S.; Sharpless, K. B. Chem.
ReV. 1994, 94, 2483. (b) Tori, S.; Liu, P.; Bhuvaneswari, N.; Amatore,
C.; Jutand, A. J. Org. Chem. 1996, 61, 3055.
Choudary, B. M.; Chowdari, N. S.; Jyothi, K.; Kantam, M. L. J. Am.
Chem. Soc. 2002, 124, 5341.
Choudary, B. M.; Chowdari, N. S.; Kantam, M. L.; Raghavan, K. V.
J. Am. Chem. Soc. 2001, 123, 9220.
(a) Schröder, M. Chem. ReV. 1980, 80, 187. (b) Wai, J. S. M.; Marko,
I.; Svendsen, J. S.; Finn, M. G.; Jacobsen, E. N.; Sharpless, K. B.
J. Am. Chem. Soc. 1989, 111, 1123. (c) Severeyns, A.; De Vos, D. E.;
Fiermans, L.; Verpoort, F.; Grobet, P. J.; Jacobs, P. A. Angew. Chem.,
Int. Ed. 2001, 40, 586.
(a) Kobayashi, S.; Ishida, T.; Akiyama, R. Org. Lett. 2001, 3, 2649.
(b) Kobayashi, S.; Endo, M.; Nagayama, S. J. Am. Chem. Soc. 1999,
121, 11229.
Mukherjee, M.; Ray, A. R. Catal. Commun. 2007, 8, 1431.
Lindsay Smith, J. R. In Metalloporphyrins in Catalytic Oxidations;
Sheldon, R. A., Ed.; Marcel Dekker: New York, 1994; Chapter 11.
Feng, K.; Wu, L. Z.; Zhang, L. P.; Tung, C. H. Tetrahedron 2007,
63, 4907.
Bonchio, M.; Carraro, M.; Scorrano, G.; Fontananova, E.; Drioli, E.
AdV. Synth. Catal. 2003, 345, 1119.
Farhadi, S.; Afshari, M. J. Chem. Res. 2006, 188.
Guo, Y.; Hu, C.; Wang, X.; Wang, Y.; Wang, E.; Zou, Y.; Ding, H.;
Feng, S. Chem. Mater. 2001, 13, 4058.
Fornal, E.; Giannotti, C. J. Photochem. Photobiol., A: Chem. 2007,
188, 279.
Kwan, C. Y.; Chu, W.; Lam, W. S. J. Mol. Catal., A: Chem. 2007,
274, 50.
Kim, S. H.; Somorjai, G. A. Proc. Natl. Acad. Sci. U.S.A. 2006, 103,
15289.
Hlatky, G. G. Chem. ReV. 2000, 100, 1347.
Li, Z.; Li, H.; Zhang, Y.; Xue, M.; Zhou, L.; Liu, Y. Appl. Catal.,
A: Gen. 2005, 292, 61.
(a) Hong, S. C.; Matyjaszewski, K. Polym. Prepr. (Am. Chem. Soc.,
DiV. Polym. Chem.) 2002, 43, 191. (b) Hong, S. C.; Matyjaszewski,
K. Macromolecules 2002, 35, 7592.
Guo, J.; Liu, B.; Wang, X.; Sun, J. React. Funct. Polym. 2004, 61,
163.
(a) Prüsse, U.; Hähnlein, M.; Daum, J.; Vorlop, K. D. Catal. Today
2000, 55, 79. (b) Centi, G.; Perathoner, S. Chem. Ind. (Milan) 2001,
83, 43. (c) Reddy, K. J.; Lin, J. Water Res. 2000, 34, 995.
Gašparovičová, D.; Králik, M.; Hronec, M.; Vallušová, Z.; Vinek,
H.; Corain, B. J. Mol. Catal., A: Chem. 2007, 264, 93.
Lakshmi Kantam, M.; Ramani, T.; Chakrapani, L.; Choudary, B. M.
J. Mol. Catal., A: Chem. 2007, 274, 11.
(a) Trevisan, V.; Signoretto, M.; Colonna, S.; Pironti, V.; Strukul,
G. Angew. Chem., Int. Ed. 2004, 43, 4097. (b) van der Made, A. W.;
Smeets, J. W. H.; Nolte, R. J. M.; Drenth, W. J. Chem. Soc., Chem.
Commun. 1983, 1204. (c) Gorman, C. B.; Parkhurst, B. L.; Su, W. L.;
Chen, K. Y. J. Am. Chem. Soc. 1997, 119, 1141. (d) Grasselli, R. K.
Top. Catal. 2001, 15, 93. (e) Cohen, B. J.; Kraus, M. A.; Patchornik,
A. J. Am. Chem. Soc. 1981, 103, 7620. (f) Voit, B. Angew. Chem.,
Int. Ed. 2006, 45, 4238. (g) Lange, M.; Martinola, F.; Oeckl, S.
Hydrocarbon Proc. 1985, 12, 51.
(a) Dorta, R.; Egli, P.; Zurcher, F.; Togni, A. J. Am. Chem. Soc.
1997, 119, 10857. (b) Lautens, M.; Fagnou, K. J. Am. Chem. Soc.
2001, 123, 7170.
(a) Smidt, S. P.; Zimmermann, N.; Studer, M.; Pfaltz, A. Chem.sEur.
J. 2004, 10, 4685. (b) Song, F.; Lancaster, S. J.; Cannon, R. D.;
Schormann, M.; Humphrey, S. M.; Zuccaccia, C.; Macchioni, A.;
Bochmann, M. Organometallics 2005, 24, 1315.
Sablong, R.; van der Vlugt, J. I.; Thomann, R.; Mecking, S.; Vogt,
D. AdV. Synth. Catal. 2005, 347, 633.
Schwab, E.; Mecking, S. Organometallics 2001, 20, 5504.
Corain, B.; Zecca, M.; Jerabek, K. J. Mol. Catal., A 2001, 177, 3.
Baek, K. W.; Park, S. W.; Nho, Y. C.; Hwang, T. S. Polymer (Korea)
2007, 31, 315.
Radkevich, V. Z.; Kistanova, I. E.; Soldatov, V. S.; Egiazarov, Y. G.
Russ. J. Appl. Chem. 2001, 74, 1446.
Morgan, N. Tech. Text. Int. 1997, 26.
Kautzmann, R. M.; Cortina, J. L.; Sampaio, C. H.; Soldatov, V. S.;
Shunkevich, A In Ion Exchange at the Millennium, Proceedings of
IEX, 2000; Cambridge, U.K.,July 16-21, 2000.
Amounas, M.; Magne, V.; Innocent, C.; Dejean, E.; Seta, P. Enzymol.
Microb. Technol. 2002, 31, 171.
Ion Exchange Resins
(208) (a) Inoue, H.; Yamanaka, K.; Yoshida, A.; Aoki, T.; Teraguchi, M.;
Kaneko, T. Polymer 2004, 45, 3. (b) Sherrington, D. C. Makromol.
Chem. Macromol. Symp. 1993, 70/71, 303.
(209) Aoki, H.; Miyano, K.; Yano, D.; Sano, K.; Yamanaka, K.; Kimura,
C.; Sugino, T. Polym. Eng. Sci. 2007, 47, 1666.
(210) Yoshimoto, N.; Nishijima, Y.; Akbarzadehlaleh, P.; Fujii, S.; Abe,
M.; Yamamoto, S. J. Chem. Eng. Jpn. 2008, 41, 200.
(211) Hutchinson, J. P.; Hilder, E. F.; Shellie, R. A.; Smith, J. A.; Haddad,
P. R. Analyst 2006, 131, 215.
(212) Aoki, H.; Miyano, K.; Hotta, S.; Yano, D.; Sano, K.; Yamanaka,
K.; Kimura, C.; Sugino, T. Electrochim. Acta 2008, 53, 6657.
(213) (a) Blanco-Brieva, G.; Cano-Serrano, E.; Campos-Martin, J. M.;
Fierro, J. L. G. Chem. Commun. 2004, 1184. (b) Sato, M.; Oono, K.
WO 2004-072,019. (c) Bombi, G.; Lora, S.; Zancato, M.; D’Archivio,
A. A.; Jerabek, K.; Corain, B. J. Mol. Catal., A: Chem. 2003, 194,
273.
(214) (a) Mennecke, K.; Cecilia, R.; Glasnov, T. N.; Gruhl, S.; Vogt, C.;
Feldhoff, A.; Vargas, M. A. L.; Kappe, C. O.; Kunz, U.; Kirschning,
Chemical Reviews, 2009, Vol. 109, No. 2 529
(215)
(216)
(217)
(218)
(219)
(220)
(221)
A. AdV. Synth. Catal. 2008, 350, 717. (b) Solodenko, W.; Wen, H.;
Leue, S.; Stuhlmann, F.; Sourkouni-Argirusi, G.; Jas, G.; Schönfeld,
H.; Kunz, U.; Kirschning, A. Eur. J. Org. Chem. 2004, 36.
Waller, F. J.; van Scoyoc, R. W. CHEMTECH 1987, 17, 438.
Harmer, M. A.; Farneth, W. E.; Sun, Q. J. Am. Chem. Soc. 1996,
118, 7708.
Harmer, M. A.; Sun, Q.; Vega, A. J.; Farneth, W. E.; Heidekum, A.;
Hölderich, W. F. Green Chem. 2000, 2, 7.
Laufer, W.; Hoelderich, W. F. Chem. Commun. 2002, 1684.
Kemmere, M. F.; Keurentjes, J. T. F. Industrial Membrane Reactors.
In Membrane Technology in the Chemical Industry; Nunes, S. P.,
Peinemann, K. V., Eds.; Wiley-VCH: Weinheim, 2001; Chapter 5.
Miachon, S.; Dalmon, J. A. Top. Catal. 2004, 29, 59.
Xu, T. J. Membr. Sci. 2005, 263, 1.
CR800404J